Hostname: page-component-76fb5796d-9pm4c Total loading time: 0 Render date: 2024-04-25T16:25:10.681Z Has data issue: false hasContentIssue false

Impact of the IVF laboratory environment on human preimplantation embryo phenotype

Part of: DOHAD & IVF

Published online by Cambridge University Press:  14 June 2017

D. K. Gardner*
Affiliation:
School of BioSciences, University of Melbourne, Parkville, VIC, Australia
R. L. Kelley
Affiliation:
School of BioSciences, University of Melbourne, Parkville, VIC, Australia
*
*Address for correspondence: D. K. Gardner, School of BioSciences, University of Melbourne, Parkville, VIC 3010, Australia. (Email david.gardner@unimelb.edu.au)

Abstract

The phenotype of the human embryo conceived through in vitro fertilization (IVF), that is its morphology, developmental kinetics, physiology and metabolism, can be affected by numerous components of the laboratory and embryo culture system (which comprise the laboratory environment). The culture media formulation is important in determining embryo phenotype, but this exists within a culture system that includes oxygen, temperature, pH and whether an embryo is cultured individually or in a group, all of which can influence embryo development. Significantly, exposure of an embryo to one suboptimal component of the culture system of laboratory typically predisposes the embryo to become more vulnerable to a second stressor, as has been well documented for atmospheric oxygen and individual culture, as well as for oxygen and ammonium. Furthermore, the inherent viability of the human embryo is derived from the quality of the gametes from which it is created. Patient age, aetiology, genetics, lifestyle (as well as ovarian stimulation in women) are all known to affect the developmental potential of gametes and hence the embryo. Thus, as well as considering the impact of the IVF laboratory environment, one needs to be aware of the status of the infertile couple, as this impacts how their gametes and embryos will respond to an in vitro environment. Although far from straight forward, analysing the interactions that exist between the human embryo and its environment will facilitate the creation of more effective and safer treatments for the infertile couple.

Type
Original Article
Copyright
© Cambridge University Press and the International Society for Developmental Origins of Health and Disease 2017 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

1. Gardner, D, Lane, M. Towards a single embryo transfer. Reprod Biomed Online. 2003; 6, 470481.CrossRefGoogle ScholarPubMed
2. Fortier, AL, McGraw, S, Lopes, FL, et al. Modulation of imprinted gene expression following superovulation. Mol Cell Endocrinol. 2014; 388, 5157.Google Scholar
3. Huffman, SR, Pak, Y, Rivera, RM. Superovulation induces alterations in the epigenome of zygotes, and results in differences in gene expression at the blastocyst stage in mice. Mol Reprod Dev. 2015; 82, 207217.Google Scholar
4. Santos, MA, Kuijk, EW, Macklon, NS. The impact of ovarian stimulation for IVF on the developing embryo. Reproduction. 2010; 139, 2334.Google Scholar
5. Wale, PL, Gardner, DK. The effects of chemical and physical factors on mammalian embryo culture and their importance for the practice of assisted human reproduction. Hum Reprod Update. 2016; 22, 222.Google Scholar
6. Finger, BJ, Harvey, AJ, Green, MP, Gardner, DK. Combined parental obesity negatively impacts preimplantation mouse embryo development, kinetics, morphology and metabolism. Hum Reprod. 2015; 30, 20842096.Google Scholar
7. Lane, M, Robker, RL, Robertson, SA. Parenting from before conception. Science. 2014; 345, 756760.Google Scholar
8. Gardner, DK, Lane, M. Ex vivo early embryo development and effects on gene expression and imprinting. Reprod Fertil Dev. 2005; 17, 361.Google Scholar
9. Gardner, DK, Larman, MG, Thouas, GA. Sex-related physiology of the preimplantation embryo. Mol Hum Reprod. 2010; 16, 539547.Google Scholar
10. Lane, M, Gardner, DK. Understanding cellular disruptions during early embryo development that perturb viability and fetal development. Reprod Fertil Dev. 2005; 17, 371378.Google Scholar
11. Wale, PL, Gardner, DK. Time-lapse analysis of mouse embryo development in oxygen gradients. Reprod Biomed Online. 2010; 21, 402410.CrossRefGoogle ScholarPubMed
12. Picton, HM, Elder, K, Houghton, FD, et al. Association between amino acid turnover and chromosome aneuploidy during human preimplantation embryo development in vitro. Mol Hum Reprod. 2010; 16, 557569.Google Scholar
13. Gardner, DK, Wale, PL, Collins, R, Lane, M. Glucose consumption of single post-compaction human embryos is predictive of embryo sex and live birth outcome. Hum Reprod. 2011; 26, 19811986.Google Scholar
14. Bermejo-Alvarez, P, Rizos, D, Lonergan, P, Gutierrez-Adan, A. Transcriptional sexual dimorphism during preimplantation embryo development and its consequences for developmental competence and adult health and disease. Reproduction. 2011; 141, 563570.Google Scholar
15. Kobayashi, S, Isotani, A, Mise, N, et al. Comparison of gene expression in male and female mouse blastocysts revealed imprinting of the X-linked gene, Rhox5/Pem, at preimplantation stages. Curr Biol. 2006; 16, 166172.Google Scholar
16. Garcia-Herreros, M, Aparicio, IM, Rath, D, Fair, T, Lonergan, P. Differential glycolytic and glycogenogenic transduction pathways in male and female bovine embryos produced in vitro. Reprod Fertil Dev. 2012; 24, 344352.Google Scholar
17. Alfarawati, S, Fragouli, E, Colls, P, et al. The relationship between blastocyst morphology, chromosomal abnormality, and embryo gender. Fertil Steril. 2011; 95, 520524.Google Scholar
18. Bodri, D, Kawachiya, S, Sugimoto, T, et al. Time-lapse variables and embryo gender: a retrospective analysis of 81 live births obtained following minimal stimulation and single embryo transfer. J Assist Reprod Genet. 2016; 33, 589596.Google Scholar
19. Bredbacka, K, Bredbacka, P. Glucose controls sex-related growth rate differences of bovine embryos produced in vitro. J Reprod Fertil. 1996; 106, 169172.Google Scholar
20. Jimenez, A, Fernandez, R, Madrid-Bury, N, et al. Experimental demonstration that pre- and post-conceptional mechanisms influence sex ratio in mouse embryos. Mol Reprod Dev. 2003; 66, 162165.CrossRefGoogle ScholarPubMed
21. Ghys, E, Dallemagne, M, De Troy, D, et al. Female bovine blastocysts are more prone to apoptosis than male ones. Theriogenology. 2016; 85, 591600.CrossRefGoogle ScholarPubMed
22. Siqueira, LG, Hansen, PJ. Sex differences in response of the bovine embryo to colony-stimulating factor 2. Reproduction. 2016; 152, 645654.Google Scholar
23. Kimura, K, Spate, LD, Green, MP, Roberts, RM. Effects of oxidative stress and inhibitors of the pentose phosphate pathway on sexually dimorphic production of IFN-tau by bovine blastocysts. Mol Reprod Dev. 2004; 68, 8895.CrossRefGoogle ScholarPubMed
24. Perez-Crespo, M, Ramirez, MA, Fernandez-Gonzalez, R, et al. Differential sensitivity of male and female mouse embryos to oxidative induced heat-stress is mediated by glucose-6-phosphate dehydrogenase gene expression. Mol Reprod Dev. 2005; 72, 502510.Google Scholar
25. Gardner, DK, Kelley, RL. Male and female embryos differ in their response to oxygen concentration. Fertil Steril. 2013; 100, S242.Google Scholar
26. Maas, K, Galkina, E, Thornton, K, Penzias, AS, Sakkas, D. No change in live birthweight of IVF singleton deliveries over an 18-year period despite significant clinical and laboratory changes. Hum Reprod. 2016; 31, 19871996.CrossRefGoogle ScholarPubMed
27. Scherrer, U, Rexhaj, E, Allemann, Y, Sartori, C, Rimoldi, SF. Cardiovascular dysfunction in children conceived by assisted reproductive technologies. Eur Heart J. 2015; 36, 15831589.Google Scholar
28. Guo, XY, Liu, XM, Jin, L, et al. Cardiovascular and metabolic profiles of offspring conceived by assisted reproductive technologies: a systematic review and meta-analysis. Fertil Steril. 2017; 107, 622631.e5.CrossRefGoogle ScholarPubMed
29. Feuer, S, Rinaudo, P. From embryos to adults: a DOHaD perspective on in vitro fertilization and other assisted reproductive technologies. Healthcare. 2016; 4, E51.Google Scholar
30. Chen, M, Heilbronn, LK. The health outcomes of human offspring conceived by assisted reproductive technologies (ART). J Dev Orig Health Dis. 2017; 115.Google ScholarPubMed
31. Mortimer, D. Quality management in the IVF laboratory. In Towards Reproductive Certainty (eds. Jansen R, Mortimer D), 1999; pp. 421426. Parthenon Publishing Group: New York.Google Scholar
32. Cummins, JM, Breen, TM, Harrison, KL, et al. A formula for scoring human embryo growth rates in in vitro fertilization: its value in predicting pregnancy and in comparison with visual estimates of embryo quality. J In Vitro Fert Embryo Transf. 1986; 3, 284295.Google Scholar
33. Edwards, RG, Fishel, SB, Cohen, J, et al. Factors influencing the success of in vitro fertilization for alleviating human infertility. J In Vitro Fert Embryo Transf. 1984; 1, 323.Google Scholar
34. Cohen, J, Trounson, A, Dawson, K, et al. The early days of IVF outside the UK. Hum Reprod Update. 2005; 11, 439459.Google Scholar
35. Scott, LA, Smith, S. The successful use of pronuclear embryo transfers the day following oocyte retrieval. Hum Reprod. 1998; 13, 10031013.Google Scholar
36. Tesarik, J, Greco, E. The probability of abnormal preimplantation development can be predicted by a single static observation on pronuclear stage morphology. Hum Reprod. 1999; 14, 13181323.CrossRefGoogle ScholarPubMed
37. Van Royen, E, Mangelschots, K, De Neubourg, D, et al. Characterization of a top quality embryo, a step towards single-embryo transfer. Hum Reprod. 1999; 14, 23452349.Google Scholar
38. Gardner, DK, Lane, M, Stevens, J, Schlenker, T, Schoolcraft, WB. Blastocyst score affects implantation and pregnancy outcome: towards a single blastocyst transfer. Fertil Steril. 2000; 73, 11551158.Google Scholar
39. Gardner, DK, Balaban, B. Assessment of human embryo development using morphological criteria in an era of time-lapse, algorithms and ‘OMICS’: is looking good still important? Mol Hum Reprod. 2016; 22, 704718.CrossRefGoogle Scholar
40. Pribenszky, C, Matyas, S, Kovacs, P, et al. Pregnancy achieved by transfer of a single blastocyst selected by time-lapse monitoring. Reprod Biomed Online. 2010; 21, 533536.Google Scholar
41. Meseguer, M, Herrero, J, Tejera, A, et al. The use of morphokinetics as a predictor of embryo implantation. Hum Reprod. 2011; 26, 26582671.Google Scholar
42. Liu, Y, Chapple, V, Roberts, P, Matson, P. Prevalence, consequence, and significance of reverse cleavage by human embryos viewed with the use of the embryoscope time-lapse video system. Fertil Steril. 2014; 102, 12951300.e2.Google Scholar
43. Motato, Y, de los Santos, MJ, Escriba, MJ, et al. Morphokinetic analysis and embryonic prediction for blastocyst formation through an integrated time-lapse system. Fertil Steril. 2016; 105, 376384.e9.Google Scholar
44. Petersen, BM, Boel, M, Montag, M, Gardner, DK. Development of a generally applicable morphokinetic algorithm capable of predicting the implantation potential of embryos transferred on day 3. Hum Reprod. 2016; 31, 22312244.Google Scholar
45. Leese, HJ, Barton, AM. Pyruvate and glucose uptake by mouse ova and preimplantation embryos. J Reprod Fertil. 1984; 72, 913.Google Scholar
46. Gardner, DK, Leese, HJ. Non-invasive measurement of nutrient uptake by single cultured pre-implantation mouse embryos. Hum Reprod. 1986; 1, 2527.Google Scholar
47. Hardy, K, Hooper, MA, Handyside, AH, et al. Non-invasive measurement of glucose and pyruvate uptake by individual human oocytes and preimplantation embryos. Hum Reprod. 1989; 4, 188191.Google Scholar
48. Houghton, FD, Hawkhead, JA, Humpherson, PG, et al. Non-invasive amino acid turnover predicts human embryo developmental capacity. Hum Reprod. 2002; 17, 9991005.CrossRefGoogle ScholarPubMed
49. Renard, JP, Philippon, A, Menezo, Y. In-vitro uptake of glucose by bovine blastocysts. J Reprod Fertil. 1980; 58, 161164.Google Scholar
50. Gardner, DK, Leese, HJ. Assessment of embryo viability prior to transfer by the noninvasive measurement of glucose uptake. J Exp Zool. 1987; 242, 103105.Google Scholar
51. Gardner, DK, Sakkas, D. Mouse embryo cleavage, metabolism and viability: role of medium composition. Hum Reprod. 1993; 8, 288295.Google Scholar
52. Lane, M, Gardner, DK. Selection of viable mouse blastocysts prior to transfer using a metabolic criterion. Hum Reprod. 1996; 11, 19751978.Google Scholar
53. Lane, M, Gardner, DK. Amino acids and vitamins prevent culture-induced metabolic perturbations and associated loss of viability of mouse blastocysts. Hum Reprod. 1998; 13, 991997.Google Scholar
54. Gardner, DK, Lane, M, Stevens, J, Schoolcraft, WB. Noninvasive assessment of human embryo nutrient consumption as a measure of developmental potential. Fertil Steril. 2001; 76, 11751180.Google Scholar
55. Tejera, A, Herrero, J, Viloria, T, et al. Time-dependent O2 consumption patterns determined optimal time ranges for selecting viable human embryos. Fertil Steril. 2012; 98, 849857.CrossRefGoogle ScholarPubMed
56. Gardner, DK. Lactate production by the mammalian blastocyst: manipulating the microenvironment for uterine implantation and invasion? Bioessays. 2015; 37, 364371.Google Scholar
57. Gardner, DK, Harvey, AJ. Blastocyst metabolism. Reprod Fertil Dev. 2015; 27, 638654.Google Scholar
58. Harvey, AJ, Rathjen, J, Gardner, DK. Metaboloepigenetic regulation of pluripotent stem cells. Stem Cells Int. 2016; 2016, 1816525.Google Scholar
59. Dominguez, F, Gadea, B, Mercader, A, et al. Embryologic outcome and secretome profile of implanted blastocysts obtained after coculture in human endometrial epithelial cells versus the sequential system. Fertil Steril. 2010; 93, 774782.e1.Google Scholar
60. Katz-Jaffe, MG, Schoolcraft, WB, Gardner, DK. Analysis of protein expression (secretome) by human and mouse preimplantation embryos. Fertil Steril. 2006; 86, 678685.Google Scholar
61. Jones, KP, Warnock, SH, Urry, RL, Edwin, SS, Mitchell, MD. Immunosuppressive activity and alpha interferon concentrations in human embryo culture media as an index of potential for successful implantation. Fertil Steril. 1992; 57, 637640.Google Scholar
62. O’Neill, C, Pike, IL, Porter, RN, et al. Maternal recognition of pregnancy prior to implantation: methods for monitoring embryonic viability in vitro and in vivo. Ann N Y Acad Sci. 1985; 442, 429439.Google Scholar
63. Smart, YC, Cripps, AW, Clancy, RL, et al. Detection of an immunosuppressive factor in human preimplantation embryo cultures. Med J Aust. 1981; 1, 7879.Google Scholar
64. Cocchiara, R, Di Trapani, G, Azzolina, A, et al. Isolation of a histamine releasing factor from human embryo culture medium after in-vitro fertilization. Hum Reprod. 1987; 2, 341344.Google Scholar
65. O’Neill, C. The role of paf in embryo physiology. Hum Reprod Update. 2005; 11, 215228.Google Scholar
66. Fuzzi, B, Rizzo, R, Criscuoli, L, et al. HLA-G expression in early embryos is a fundamental prerequisite for the obtainment of pregnancy. Eur J Immunol. 2002; 32, 311315.Google Scholar
67. Sher, G, Keskintepe, L, Nouriani, M, Roussev, R, Batzofin, J. Expression of sHLA-G in supernatants of individually cultured 46-h embryos: a potentially valuable indicator of ‘embryo competency’ and IVF outcome. Reprod Biomed Online. 2004; 9, 7478.Google Scholar
68. Noci, I, Fuzzi, B, Rizzo, R, et al. Embryonic soluble HLA-G as a marker of developmental potential in embryos. Hum Reprod. 2005; 20, 138146.Google Scholar
69. Kotze, D, Kruger, TF, Lombard, C, et al. The effect of the biochemical marker soluble human leukocyte antigen G on pregnancy outcome in assisted reproductive technology – a multicenter study. Fertil Steril. 2013; 100, 13031309.Google Scholar
70. Vercammen, MJ, Verloes, A, Van de Velde, H, Haentjens, P. Accuracy of soluble human leukocyte antigen-G for predicting pregnancy among women undergoing infertility treatment: meta-analysis. Hum Reprod Update. 2008; 14, 209218.Google Scholar
71. Thouas, GA, Dominguez, F, Green, MP, et al. Soluble ligands and their receptors in human embryo development and implantation. Endocr Rev. 2015; 36, 92130.Google Scholar
72. Sheth, KV, Roca, GL, al-Sedairy, ST, et al. Prediction of successful embryo implantation by measuring interleukin-1-alpha and immunosuppressive factor(s) in preimplantation embryo culture fluid. Fertil Steril. 1991; 55, 952957.Google Scholar
73. Sequeira, K, Espejel-Nunez, A, Vega-Hernandez, E, Molina-Hernandez, A, Grether-Gonzalez, P. An increase in IL-1beta concentrations in embryo culture-conditioned media obtained by in vitro fertilization on day 3 is related to successful implantation. J Assist Reprod Genet. 2015; 32, 16231627.Google Scholar
74. Katz-Jaffe, MG, Gardner, DK, Schoolcraft, WB. Proteomic analysis of individual human embryos to identify novel biomarkers of development and viability. Fertil Steril. 2006; 85, 101107.Google Scholar
75. McReynolds, S, Vanderlinden, L, Stevens, J, et al. Lipocalin-1: a potential marker for noninvasive aneuploidy screening. Fertil Steril. 2011; 95, 26312633.Google Scholar
76. Nyalwidhe, J, Burch, T, Bocca, S, et al. The search for biomarkers of human embryo developmental potential in IVF: a comprehensive proteomic approach. Mol Hum Reprod. 2013; 19, 250263.Google Scholar
77. Dominguez, F, Meseguer, M, Aparicio-Ruiz, B, et al. New strategy for diagnosing embryo implantation potential by combining proteomics and time-lapse technologies. Fertil Steril. 2015; 104, 908914.Google Scholar
78. Huang, G, Zhou, C, Wei, CJ, et al. Evaluation of in vitro fertilization outcomes using interleukin-8 in culture medium of human preimplantation embryos. Fertil Steril. 2017; 107, 649656.Google Scholar
79. Montsko, G, Zrinyi, Z, Janaky, T, et al. Noninvasive embryo viability assessment by quantitation of human haptoglobin alpha-1 fragment in the in vitro fertilization culture medium: an additional tool to increase success rate. Fertil Steril. 2015; 103, 687693.Google Scholar
80. Cortezzi, SS, Garcia, JS, Ferreira, CR, et al. Secretome of the preimplantation human embryo by bottom-up label-free proteomics. Anal Bioanal Chem. 2011; 401, 13311339.Google Scholar
81. Kropp, J, Salih, SM, Khatib, H. Expression of microRNAs in bovine and human pre-implantation embryo culture media. Front Genet. 2014; 5, 91.Google Scholar
82. Rosenbluth, EM, Shelton, DN, Wells, LM, Sparks, AE, Van Voorhis, BJ. Human embryos secrete microRNAs into culture media – a potential biomarker for implantation. Fertil Steril. 2014; 101, 14931500.Google Scholar
83. Cuman, C, Van Sinderen, M, Gantier, MP, et al. Human blastocyst secreted microRNA regulate endometrial epithelial cell adhesion. EBioMedicine. 2015; 2, 15281535.Google Scholar
84. Capalbo, A, Ubaldi, FM, Cimadomo, D, et al. MicroRNAs in spent blastocyst culture medium are derived from trophectoderm cells and can be explored for human embryo reproductive competence assessment. Fertil Steril. 2016; 105, 225235.e1–3.CrossRefGoogle ScholarPubMed
85. Stigliani, S, Anserini, P, Venturini, PL, Scaruffi, P. Mitochondrial DNA content in embryo culture medium is significantly associated with human embryo fragmentation. Hum Reprod. 2013; 28, 26522660.Google Scholar
86. Galluzzi, L, Palini, S, Stefani, SD, et al. Extracellular embryo genomic DNA and its potential for genotyping applications. Future Sci OA. 2015; 1, FSO62.Google Scholar
87. Shamonki, MI, Jin, H, Haimowitz, Z, Liu, L. Proof of concept: preimplantation genetic screening without embryo biopsy through analysis of cell-free DNA in spent embryo culture media. Fertil Steril. 2016; 106, 13121318.Google Scholar
88. Xu, J, Fang, R, Chen, L, et al. Noninvasive chromosome screening of human embryos by genome sequencing of embryo culture medium for in vitro fertilization. Proc Natl Acad Sci U S A. 2016; 113, 1190711912.Google Scholar
89. Lee, YS, Thouas, GA, Gardner, DK. Developmental kinetics of cleavage stage mouse embryos are related to their subsequent carbohydrate and amino acid utilization at the blastocyst stage. Hum Reprod. 2015; 30, 543552.CrossRefGoogle ScholarPubMed
90. Tejera, A, Castello, D, de Los Santos, JM, et al. Combination of metabolism measurement and a time-lapse system provides an embryo selection method based on oxygen uptake and chronology of cytokinesis timing. Fertil Steril. 2016; 106, 119126.e2.Google Scholar
91. Milazzotto, MP, Goissis, MD, Chitwood, JL, et al. Early cleavages influence the molecular and the metabolic pattern of individually cultured bovine blastocysts. Mol Reprod Dev. 2016; 83, 324336.Google Scholar
92. Fragouli, E, Lenzi, M, Ross, R, et al. Comprehensive molecular cytogenetic analysis of the human blastocyst stage. Hum Reprod. 2008; 23, 25962608.Google Scholar
93. Munne, S, Magli, C, Adler, A, et al. Treatment-related chromosome abnormalities in human embryos. Hum Reprod. 1997; 12, 780784.Google Scholar
94. Bavister, BD. Culture of preimplantation embryos: facts and artifacts. Hum Reprod Update. 1995; 1, 91148.Google Scholar
95. Biggers, JD, Borland, RM. Physiological aspects of growth and development of the preimplantation mammalian embryo. Annu Rev Physiol. 1976; 38, 95119.Google Scholar
96. Leese, HJ, Donnay, I, Thompson, JG. Human assisted conception: a cautionary tale. Lessons from domestic animals. Hum Reprod. 1998; 13(Suppl. 4), 184202.Google Scholar
97. Gardner, DK. Mammalian embryo culture in the absence of serum or somatic cell support. Cell Biol Int. 1994; 18, 11631179.Google Scholar
98. Gardner, DK, Lane, M. Embryo culture systems. In In Vitro Fertilization: A Practical Approach (ed. Gardner DK), 2007; pp. 221282. Informa Healthcare: New York.Google Scholar
99. Basile, N, Morbeck, D, Garcia-Velasco, J, Bronet, F, Meseguer, M. Type of culture media does not affect embryo kinetics: a time-lapse analysis of sibling oocytes. Hum Reprod. 2013; 28, 634641.Google Scholar
100. Ciray, HN, Aksoy, T, Goktas, C, Ozturk, B, Bahceci, M. Time-lapse evaluation of human embryo development in single versus sequential culture media – a sibling oocyte study. J Assist Reprod Genet. 2012; 29, 891900.Google Scholar
101. Hickman, C, Wells, D, Lavery, S. Laboratory culture environment affects mitotic aneuploidy, not meiotic anueploidy. 2017; pp. 143. Proceedings of the British Fertility Society: Edinburgh.Google Scholar
102. Edwards, LJ, Williams, DA, Gardner, DK. Intracellular pH of the mouse preimplantation embryo: amino acids act as buffers of intracellular pH. Hum Reprod. 1998; 13, 34413448.Google Scholar
103. Lane, M. Mechanisms for managing cellular and homeostatic stress in vitro. Theriogenology. 2001; 55, 225236.Google Scholar
104. Dawson, KM, Baltz, JM. Organic osmolytes and embryos: substrates of the Gly and beta transport systems protect mouse zygotes against the effects of raised osmolarity. Biol Reprod. 1997; 56, 15501558.Google Scholar
105. Lane, M, Gardner, DK. Mitochondrial malate-aspartate shuttle regulates mouse embryo nutrient consumption. J Biol Chem. 2005; 280, 1836118367.Google Scholar
106. Martin, PM, Sutherland, AE. Exogenous amino acids regulate trophectoderm differentiation in the mouse blastocyst through an mTOR-dependent pathway. Dev Biol. 2001; 240, 182193.Google Scholar
107. Tan, BS, Rathjen, PD, Harvey, AJ, Gardner, DK, Rathjen, J. Regulation of amino acid transporters in pluripotent cell populations in the embryo and in culture; novel roles for sodium-coupled neutral amino acid transporters. Mech Dev. 2016; 141, 3239.Google Scholar
108. Gardner, DK. Dissection of culture media for embryos: the most important and less important components and characteristics. Reprod Fertil Dev. 2008; 20, 918.Google Scholar
109. Lane, M, Gardner, DK. Embryo culture medium: which is the best? Best Pract Res Clin Obstet Gynaecol. 2007; 21, 83100.Google Scholar
110. Gardner, DK, Lane, M. Amino acids and ammonium regulate mouse embryo development in culture. Biol Reprod. 1993; 48, 377385.Google Scholar
111. Devreker, F, Hardy, K, Van den Bergh, M, et al. Amino acids promote human blastocyst development in vitro. Hum Reprod. 2001; 16, 749756.Google Scholar
112. Heeneman, S, Deutz, NE, Buurman, WA. The concentrations of glutamine and ammonia in commercially available cell culture media. J Immunol Methods. 1993; 166, 8591.CrossRefGoogle ScholarPubMed
113. Gardner, DK, Lane, M, Spitzer, A, Batt, PA. Enhanced rates of cleavage and development for sheep zygotes cultured to the blastocyst stage in vitro in the absence of serum and somatic cells: amino acids, vitamins, and culturing embryos in groups stimulate development. Biol Reprod. 1994; 50, 390400.Google Scholar
114. Lane, M, Gardner, DK. Culture of preimplantation mouse embryos in the presence of amino acids increases post-implantation development whilst the concomitant build up of ammonium induces birth defects. J Reprod Fertil. 1994; 102, 305312.Google Scholar
115. Zander, DL, Froiland, D, Lane, M. Ammonium effects mitochondrial distribution and function in mouse 2-cell embryo development. Reprod Fertil Dev. 2004; 16, 81.Google Scholar
116. Lane, M, Gardner, DK. Ammonium induces aberrant blastocyst differentiation, metabolism, pH regulation, gene expression and subsequently alters fetal development in the mouse. Biol Reprod. 2003; 69, 11091117.CrossRefGoogle ScholarPubMed
117. Gardner, DK, Hamilton, R, McCallie, B, Schoolcraft, WB, Katz-Jaffe, MG. Human and mouse embryonic development, metabolism and gene expression are altered by an ammonium gradient in vitro. Reproduction. 2013; 146, 4961.Google Scholar
118. Lane, M, Gardner, DK. Increase in postimplantation development of cultured mouse embryos by amino acids and induction of fetal retardation and exencephaly by ammonium ions. J Reprod Fertil. 1994; 102, 305312.Google Scholar
119. Sinawat, S, Hsaio, WC, Flockhart, JH, et al. Fetal abnormalities produced after preimplantation exposure of mouse embryos to ammonium chloride. Hum Reprod. 2003; 18, 21572165.Google Scholar
120. Biggers, JD, McGinnis, LK, Summers, MC. Discrepancies between the effects of glutamine in cultures of preimplantation mouse embryos. Reprod Biomed Online. 2004; 9, 7073.Google Scholar
121. Wale, PL, Gardner, DK. Oxygen affects the ability of mouse blastocysts to regulate ammonium. Biol Reprod. 2013; 89, 75.Google Scholar
122. Virant-Klun, I, Tomazevic, T, Vrtacnik-Bokal, E, Vogler, A, Krsnik, M, Meden-Vrtovec, H. Increased ammonium in culture medium reduces the development of human embryos to the blastocyst stage. Fertil Steril. 2006; 85, 526528.Google Scholar
123. Hardarson, T, Bungum, M, Conaghan, J, et al. Noninferiority, randomized, controlled trial comparing embryo development using media developed for sequential or undisturbed culture in a time-lapse setup. Fertil Steril. 2015; 104, 14521459.e1–4.Google Scholar
124. Gardner, DK, Lane, M, Calderon, I, Leeton, J. Environment of the preimplantation human embryo in vivo: metabolite analysis of oviduct and uterine fluids and metabolism of cumulus cells. Fertil Steril. 1996; 65, 349353.Google Scholar
125. Biggers, JD, Whittingham, DG, Donahue, RP. The pattern of energy metabolism in the mouse oocyte and zygote. Proc Natl Acad Sci U S A. 1967; 58, 560567.Google Scholar
126. Gardner, DK, Wale, PL. Analysis of metabolism to select viable human embryos for transfer. Fertil Steril. 2013; 99, 10621072.Google Scholar
127. Gardner, DK, Leese, HJ. The role of glucose and pyruvate transport in regulating nutrient utilization by preimplantation mouse embryos. Development. 1988; 104, 423429.Google Scholar
128. Gardner, DK. Changes in requirements and utilization of nutrients during mammalian preimplantation embryo development and their significance in embryo culture. Theriogenology. 1998; 49, 83102.Google Scholar
129. Lane, M, Gardner, DK. Lactate regulates pyruvate uptake and metabolism in the preimplantation mouse embryo. Biol Reprod. 2000; 62, 1622.Google Scholar
130. Nikiforov, A, Kulikova, V, Ziegler, M. The human NAD metabolome: functions, metabolism and compartmentalization. Crit Rev Biochem Mol Biol. 2015; 50, 284297.Google Scholar
131. Ryall, JG, Dell’Orso, S, Derfoul, A, et al. The NAD(+)-dependent SIRT1 deacetylase translates a metabolic switch into regulatory epigenetics in skeletal muscle stem cells. Cell Stem Cell. 2015; 16, 171183.Google Scholar
132. Imai, S, Guarente, L. NAD+ and sirtuins in aging and disease. Trends Cell Biol. 2014; 24, 464471.CrossRefGoogle ScholarPubMed
133. Quenet, D, El Ramy, R, Schreiber, V, Dantzer, F. The role of poly(ADP-ribosyl)ation in epigenetic events. Int J Biochem Cell Biol. 2009; 41, 6065.Google Scholar
134. Canto, C, Sauve, AA, Bai, P. Crosstalk between poly(ADP-ribose) polymerase and sirtuin enzymes. Mol Aspects Med. 2013; 34, 11681201.Google Scholar
135. Conaghan, J, Handyside, AH, Winston, RM, Leese, HJ. Effects of pyruvate and glucose on the development of human preimplantation embryos in vitro. J Reprod Fertil. 1993; 99, 8795.Google Scholar
136. Xu, J, Sinclair, KD. One-carbon metabolism and epigenetic regulation of embryo development. Reprod Fertil Dev. 2015; 27, 667676.Google Scholar
137. Meldrum, DR, Chetkowski, R, Steingold, KA, et al. Evolution of a highly successful in vitro fertilization-embryo transfer program. Fertil Steril. 1987; 48, 8693.Google Scholar
138. Barnes, D, Sato, G. Methods for growth of cultured cells in serum-free medium. Anal Biochem. 1980; 102, 255270.CrossRefGoogle ScholarPubMed
139. Barnes, D, Sato, G. Serum-free cell culture: a unifying approach. Cell. 1980; 22, 649655.Google Scholar
140. Damewood, MD, Hesla, JS, Schlaff, WD, et al. Effect of serum from patients with minimal to mild endometriosis on mouse embryo development in vitro. Fertil Steril. 1990; 54, 917920.Google Scholar
141. Abu-Musa, A, Takahashi, K, Kitao, M. The effect of serum obtained before and after treatment for endometriosis on in vitro development of two-cell mouse embryos. Fertil Steril. 1992; 57, 10981102.Google Scholar
142. Dokras, A, Sargent, IL, Redman, CW, Barlow, DH. Sera from women with unexplained infertility inhibit both mouse and human embryo growth in vitro. Fertil Steril. 1993; 60, 285292.Google Scholar
143. Miller, KA, Pittaway, DE, Deaton, JL. The effect of serum from infertile women with endometriosis on fertilization and early embryonic development in a murine in vitro fertilization model. Fertil Steril. 1995; 64, 623626.Google Scholar
144. Deaton, JL, Dempsey, RA, Miller, KA. Serum from women with polycystic ovary syndrome inhibits fertilization and embryonic development in the murine in vitro fertilization model. Fertil Steril. 1996; 65, 12241228.Google Scholar
145. Menezo, Y, Testart, J, Perrone, D. Serum is not necessary in human in vitro fertilization, early embryo culture, and transfer. Fertil Steril. 1984; 42, 750755.Google Scholar
146. Leese, HJ. The formation and function of oviduct fluid. J Reprod Fertil. 1988; 82, 843856.Google Scholar
147. Khosla, S, Dean, W, Brown, D, Reik, W, Feil, R. Culture of preimplantation mouse embryos affects fetal development and the expression of imprinted genes. Biol Reprod. 2001; 64, 918926.Google Scholar
148. Fernandez-Gonzalez, R, Moreira, P, Bilbao, A, et al. Long-term effect of in vitro culture of mouse embryos with serum on mRNA expression of imprinting genes, development, and behavior. Proc Natl Acad Sci U S A. 2004; 101, 58805885.Google Scholar
149. Thompson, JG, Gardner, DK, Pugh, PA, McMillan, WH, Tervit, HR. Lamb birth weight is affected by culture system utilized during in vitro pre-elongation development of ovine embryos. Biol Reprod. 1995; 53, 13851391.Google Scholar
150. Dorland, M, Gardner, DK, Trounson, A. Serum in synthetic oviduct fluid causes mitochondrial degeneration in ovine embryos. Reprod Fertil (Abstr Ser). 1994; 13, 70.Google Scholar
151. Walker, SK, Heard, TM, Seamark, RF. In vitro culture of sheep embryos without co-culture: success and perspectives. Theriogenology. 1992; 37, 111126.Google Scholar
152. Bongso, A, Soon-Chye, N, Sathananthan, H, et al. Improved quality of human embryos when co-cultured with human ampullary cells. Hum Reprod. 1989; 4, 706713.Google Scholar
153. Ashwood-Smith, MJ, Hollands, P, Edwards, RG. The use of albuminar 5 as a medium supplement in clinical IVF. Hum Reprod. 1989; 4, 702705.Google Scholar
154. Khan, I, Staessen, C, Devroey, P, Van Steirteghem, AC. Human serum albumin versus serum: a comparative study on embryo transfer medium. Fertil Steril. 1991; 56, 98101.Google Scholar
155. Staessen, C, Van den Abbeel, E, Carle, M, et al. Comparison between human serum and albuminar-20 (TM) supplement for in-vitro fertilization. Hum Reprod. 1990; 5, 336341.Google Scholar
156. Huisman, GJ, Lo-A-Njoe, NM, Alberda, AT, et al. Comparison of results obtained with human serum and a protein solution as a supplement for in vitro fertilization culture medium. Fertil Steril. 1992; 58, 637639.Google Scholar
157. Adler, A, Reing, AM, Bedford, JM, Alikani, M, Cohen, J. Plasmanate as a medium supplement for in vitro fertilization. J Assist Reprod Genet. 1993; 10, 6771.Google Scholar
158. Pool, TB, Martin, JE. High continuing pregnancy rates after in vitro fertilization-embryo transfer using medium supplemented with a plasma protein fraction containing alpha- and beta-globulins. Fertil Steril. 1994; 61, 714719.Google Scholar
159. Weathersbee, PS, Pool, TB, Ord, T. Synthetic serum substitute (SSS): a globulin-enriched protein supplement for human embryo culture. J Assist Reprod Genet. 1995; 12, 354360.Google Scholar
160. Desai, NN, Sheean, LA, Martin, D, et al. Clinical experience with synthetic serum substitute as a protein supplement in IVF culture media: a retrospective study. J Assist Reprod Genet. 1996; 13, 2331.Google Scholar
161. Tucker, KE, Hurst, BS, Guadagnoli, S, et al. Evaluation of synthetic serum substitute versus serum as protein supplementation for mouse and human embryo culture. J Assist Reprod Genet. 1996; 13, 3237.Google Scholar
162. Meintjes, M, Chantilis, SJ, Ward, DC, et al. A randomized controlled study of human serum albumin and serum substitute supplement as protein supplements for IVF culture and the effect on live birth rates. Hum Reprod. 2009; 24, 782789.Google Scholar
163. Hanson, RW, Ballard, FJ. Citrate, pyruvate, and lactate contaminants of commercial serum albumin. J Lipid Res. 1968; 9, 667668.Google Scholar
164. Dyrlund, TF, Kirkegaard, K, Poulsen, ET, et al. Unconditioned commercial embryo culture media contain a large variety of non-declared proteins: a comprehensive proteomics analysis. Hum Reprod. 2014; 29, 24212430.Google Scholar
165. Morbeck, DE, Paczkowski, M, Fredrickson, JR, et al. Composition of protein supplements used for human embryo culture. J Assist Reprod Genet. 2014; 31, 17031711.Google Scholar
166. Batt, PA, Gardner, DK, Cameron, AW. Oxygen concentration and protein source affect the development of preimplantation goat embryos in vitro. Reprod Fertil Dev. 1991; 3, 601607.Google Scholar
167. McKiernan, SH, Bavister, BD. Different lots of bovine serum albumin inhibit or stimulate in vitro development of hamster embryos. In Vitro Cell Dev Biol. 1992; 28A, 154156.Google Scholar
168. Kane, MT. Variability in different lots of commercial bovine serum albumin affects cell multiplication and hatching of rabbit blastocysts in culture. J Reprod Fertil. 1983; 69, 555558.Google Scholar
169. Leonard, PH, Charlesworth, MC, Benson, L, et al. Variability in protein quality used for embryo culture: embryotoxicity of the stabilizer octanoic acid. Fertil Steril. 2013; 100, 544549.Google Scholar
170. Takatori, S, Akutsu, K, Kondo, F, et al. Di(2-ethylhexyl)phthalate and mono(2-ethylhexyl)phthalate in media for in vitro fertilization. Chemosphere. 2012; 86, 454459.Google Scholar
171. Akutsu, K, Takatori, S, Nakazawa, H, Makino, T. Detection of polybrominated diphenyl ethers in culture media and protein sources used for human in vitro fertilization. Chemosphere. 2013; 92, 864869.Google Scholar
172. Christianson, MS, Zhao, Y, Shoham, G, et al. Embryo catheter loading and embryo culture techniques: results of a worldwide web-based survey. J Assist Reprod Genet. 2014; 31, 10291036.Google Scholar
173. Fischer, B, Bavister, BD. Oxygen tension in the oviduct and uterus of rhesus monkeys, hamsters and rabbits. J Reprod Fertil. 1993; 99, 673679.Google Scholar
174. Mastroianni, L Jr, Jones, R. Oxygen tension within the rabbit fallopian tube. J Reprod Fertil. 1965; 147, 99102.Google Scholar
175. Ross, RN, Graves, CN. O2 levels in female rabbit reproductive tract. J Anim Sci. 1974; 39, 994.Google Scholar
176. Maas, DH, Storey, BT, Mastroianni, L Jr. Oxygen tension in the oviduct of the rhesus monkey (Macaca mulatta). Fertil Steril. 1976; 27, 13121317.Google Scholar
177. Quinn, P, Harlow, GM. The effect of oxygen on the development of preimplantation mouse embryos in vitro. J Exp Zool. 1978; 206, 7380.Google Scholar
178. Tervit, HR, Whittingham, DG, Rowson, LE. Successful culture in vitro of sheep and cattle ova. J Reprod Fertil. 1972; 30, 493497.Google Scholar
179. Whitten, WK. Nutrient requirements for the culture of preimplantation embryos in vitro. Adv Biosci. 1971; 6, 129139.Google Scholar
180. Thompson, JG, Simpson, AC, Pugh, PA, Donnelly, PE, Tervit, HR. Effect of oxygen concentration on in-vitro development of preimplantation sheep and cattle embryos. J Reprod Fertil. 1990; 89, 573578.Google Scholar
181. Kirkegaard, K, Hindkjaer, JJ, Ingerslev, HJ. Effect of oxygen concentration on human embryo development evaluated by time-lapse monitoring. Fertil Steril. 2013; 99, 738744.e4.Google Scholar
182. Harvey, AJ, Kind, KL, Pantaleon, M, Armstrong, DT, Thompson, JG. Oxygen-regulated gene expression in bovine blastocysts. Biol Reprod. 2004; 71, 11081119.Google Scholar
183. Kind, KL, Collett, RA, Harvey, AJ, Thompson, JG. Oxygen-regulated expression of GLUT-1, GLUT-3, and VEGF in the mouse blastocyst. Mol Reprod Dev. 2005; 70, 3744.Google Scholar
184. Rinaudo, PF, Giritharan, G, Talbi, S, Dobson, AT, Schultz, RM. Effects of oxygen tension on gene expression in preimplantation mouse embryos. Fertil Steril. 2006; 86, 12521265, 1265.e1–36.Google Scholar
185. Li, W, Goossens, K, Van Poucke, M, et al. High oxygen tension increases global methylation in bovine 4-cell embryos and blastocysts but does not affect general retrotransposon expression. Reprod Fertil Dev. 2016; 28, 948959.Google Scholar
186. Gaspar, RC, Arnold, DR, Correa, CA, et al. Oxygen tension affects histone remodeling of in vitro-produced embryos in a bovine model. Theriogenology. 2015; 83, 14081415.Google Scholar
187. Ghosh, J, Coutifaris, C, Sapienza, C, Mainigi, M. Global DNA methylation levels are altered by modifiable clinical manipulations in assisted reproductive technologies. Clin Epigenetics. 2017; 9, 14.Google Scholar
188. Katz-Jaffe, MG, Linck, DW, Schoolcraft, WB, Gardner, DK. A proteomic analysis of mammalian preimplantation embryonic development. Reproduction. 2005; 130, 899905.Google Scholar
189. Khurana, NK, Wales, RG. Effects of oxygen concentration on the metabolism of [U-14C]glucose by mouse morulae and early blastocysts in vitro. Reprod Fertil Dev. 1989; 1, 99106.Google Scholar
190. Wale, PL, Gardner, DK. Oxygen regulates amino acid turnover and carbohydrate uptake during the preimplantation period of mouse embryo development. Biol Reprod. 2012; 87, 24.Google Scholar
191. Dumoulin, JC, Meijers, CJ, Bras, M, et al. Effect of oxygen concentration on human in-vitro fertilization and embryo culture. Hum Reprod. 1999; 14, 465469.Google Scholar
192. Waldenstrom, U, Engstrom, AB, Hellberg, D, Nilsson, S. Low-oxygen compared with high-oxygen atmosphere in blastocyst culture, a prospective randomized study. Fertil Steril. 2009; 91, 24612465.Google Scholar
193. Kovacic, B, Vlaisavljevic, V. Influence of atmospheric versus reduced oxygen concentration on development of human blastocysts in vitro: a prospective study on sibling oocytes. Reprod Biomed Online. 2008; 17, 229236.Google Scholar
194. Mantikou, E, Jonker, MJ, Wong, KM, et al. Factors affecting the gene expression of in vitro cultured human preimplantation embryos. Hum Reprod. 2016; 31, 298311.Google Scholar
195. Catt, JW, Henman, M. Toxic effects of oxygen on human embryo development. Hum Reprod. 2000; 15(Suppl. 2), 199206.Google Scholar
196. Meintjes, M, Chantilis, SJ, Douglas, JD, et al. A controlled randomized trial evaluating the effect of lowered incubator oxygen tension on live births in a predominantly blastocyst transfer program. Hum Reprod. 2009; 24, 300307.Google Scholar
197. Gomes Sobrinho, DB, Oliveira, JB, Petersen, CG, et al. IVF/ICSI outcomes after culture of human embryos at low oxygen tension: a meta-analysis. Reprod Biol Endocrinol. 2011; 9, 143.Google Scholar
198. Kelley, RL, Gardner, DK. Combined effects of individual culture and atmospheric oxygen on preimplantation mouse embryos in vitro. Reprod Biomed Online. 2016; 33, 537549.Google Scholar
199. Kelley, RL, Gardner, DK. In vitro culture of individual mouse preimplantation embryos: the role of embryo density, microwells, oxygen, timing and conditioned media. Reprod Biomed Online. 2017; 34, 441454.Google Scholar
200. Brison, DR, Schultz, RM. Apoptosis during mouse blastocyst formation: evidence for a role for survival factors including transforming growth factor alpha. Biol Reprod. 1997; 56, 10881096.Google Scholar
201. Paria, BC, Dey, SK. Preimplantation embryo development in vitro: cooperative interactions among embryos and role of growth factors. Proc Natl Acad Sci U S A. 1990; 87, 47564760.Google Scholar
202. Lane, M, Gardner, DK. Effect of incubation volume and embryo density on the development and viability of mouse embryos in vitro. Hum Reprod. 1992; 7, 558562.Google Scholar
203. Hughes, PM, Hudson, SB, Walker, DL, et al. Mouse embryo assay (MEA): group versus individual culture effects on the sensitivity for peroxides in mineral oil. Fertil Steril. 2009; 92, S231.Google Scholar
204. Spyropoulou, I, Karamalegos, C, Bolton, VN. A prospective randomized study comparing the outcome of in-vitro fertilization and embryo transfer following culture of human embryos individually or in groups before embryo transfer on day 2. Hum Reprod. 1999; 14, 7679.Google Scholar
205. Rijnders, PM, Jansen, CA. Influence of group culture and culture volume on the formation of human blastocysts: a prospective randomized study. Hum Reprod. 1999; 14, 23332337.Google Scholar
206. Almagor, M, Bejar, C, Kafka, I, Yaffe, H. Pregnancy rates after communal growth of preimplantation human embryos in vitro. Fertil Steril. 1996; 66, 394397.Google Scholar
207. Moessner, J, Dodson, WC. The quality of human embryo growth is improved when embryos are cultured in groups rather than separately. Fertil Steril. 1995; 64, 10341035.Google Scholar
208. Ebner, T, Shebl, O, Moser, M, et al. Group culture of human zygotes is superior to individual culture in terms of blastulation, implantation and life birth. Reprod Biomed Online. 2010; 21, 762768.Google Scholar
209. Rebollar-Lazaro, I, Matson, P. The culture of human cleavage stage embryos alone or in groups: effect upon blastocyst utilization rates and implantation. Reprod Biol. 2010; 10, 227234.Google Scholar
210. Bolton, VN, Cutting, R, Clarke, H, Brison, DR. ACE consensus meeting report: culture systems. Hum Fertil (Camb). 2014; 17, 239251.Google Scholar
211. Minasi, MG, Fabozzi, G, Casciani, V, et al. Improved blastocyst formation with reduced culture volume: comparison of three different culture conditions on 1128 sibling human zygotes. J Assist Reprod Genet. 2015; 32, 215220.CrossRefGoogle ScholarPubMed
212. De Munck, N, Santos-Ribeiro, S, Mateizel, I, Verheyen, G. Reduced blastocyst formation in reduced culture volume. J Assist Reprod Genet. 2015; 32, 13651370.Google Scholar
213. Kato, Y, Tsunoda, Y. Effects of the culture density of mouse zygotes on the development in vitro and in vivo. Theriogenology. 1994; 41, 13151322.Google Scholar
214. Nagao, Y, Iijima, R, Saeki, K. Interaction between embryos and culture conditions during in vitro development of bovine early embryos. Zygote. 2008; 16, 127133.Google Scholar
215. Wiley, LM, Yamami, S, Van Muyden, D. Effect of potassium concentration, type of protein supplement, and embryo density on mouse preimplantation development in vitro. Fertil Steril. 1986; 45, 111119.Google Scholar
216. Vajta, G, Peura, TT, Holm, P, et al. New method for culture of zona-included or zona-free embryos: the well of the well (WOW) system. Mol Reprod Dev. 2000; 55, 256264.Google Scholar
217. Hoelker, M, Rings, F, Lund, Q, et al. Effect of embryo density on in vitro developmental characteristics of bovine preimplantative embryos with respect to micro and macroenvironments. Reprod Domest Anim. 2010; 45, e138e145.Google Scholar
218. Vajta, G, Korosi, T, Du, Y, et al. The well-of-the-well system: an efficient approach to improve embryo development. Reprod Biomed Online. 2008; 17, 7381.Google Scholar
219. Hashimoto, S, Kato, N, Saeki, K, Morimoto, Y. Selection of high-potential embryos by culture in poly(dimethylsiloxane) microwells and time-lapse imaging. Fertil Steril. 2012; 97, 332337.Google Scholar
220. Kahraman, S. Comparison of blastocyst development and cycle outcome in patients with eSET using either conventional or time lapse incubators. J Reprod Biotech Fertil. 2012; 3, 55.Google Scholar
221. Rubio, I, Galan, A, Larreategui, Z, et al. Clinical validation of embryo culture and selection by morphokinetic analysis: a randomized, controlled trial of the EmbryoScope. Fertil Steril. 2014; 102, 12871294.Google Scholar
222. Macklon, NS, Pieters, MH, Hassan, MA, et al. A prospective randomized comparison of sequential versus monoculture systems for in-vitro human blastocyst development. Hum Reprod. 2002; 17, 27002705.Google Scholar
223. Costa-Borges, N, Belles, M, Meseguer, M, et al. Blastocyst development in single medium with or without renewal on day 3: a prospective cohort study on sibling donor oocytes in a time-lapse incubator. Fertil Steril. 2016; 105, 707713.Google Scholar
224. Zhang, JQ, Li, XL, Peng, Y, et al. Reduction in exposure of human embryos outside the incubator enhances embryo quality and blastulation rate. Reprod Biomed Online. 2010; 20, 510515.Google Scholar
225. Heitmann, RJ, Hill, MJ, James, AN, et al. Live births achieved via IVF are increased by improvements in air quality and laboratory environment. Reprod Biomed Online. 2015; 31, 364371.Google Scholar
226. Khoudja, RY, Xu, Y, Li, T, Zhou, C. Better IVF outcomes following improvements in laboratory air quality. J Assist Reprod Genet. 2013; 30, 6976.Google Scholar
227. Boone, WR, Johnson, JE, Locke, AJ, Crane, MM, Price, TM. Control of air quality in an assisted reproductive technology laboratory. Fertil Steril. 1999; 71, 150154.Google Scholar
228. Munch, EM, Sparks, AE, Duran, HE, Van Voorhis, BJ. Lack of carbon air filtration impacts early embryo development. J Assist Reprod Genet. 2015; 32, 10091017.Google Scholar
229. Higdon, HL 3rd, Blackhurst, DW, Boone, WR. Incubator management in an assisted reproductive technology laboratory. Fertil Steril. 2008; 89, 703710.Google Scholar
230. Hall, J, Gilligan, A, Schimmel, T, Cecchi, M, Cohen, J. The origin, effects and control of air pollution in laboratories used for human embryo culture. Hum Reprod. 1998; 13(Suppl. 4), 146155.Google Scholar
231. Sifer, C, Pont, JC, Porcher, R, et al. A prospective randomized study to compare four different mineral oils used to culture human embryos in IVF/ICSI treatments. Eur J Obstet Gynecol Reprod Biol. 2009; 147, 5256.Google Scholar
232. Otsuki, J, Nagai, Y, Chiba, K. Peroxidation of mineral oil used in droplet culture is detrimental to fertilization and embryo development. Fertil Steril. 2007; 88, 741743.Google Scholar
233. Morbeck, DE, Khan, Z, Barnidge, DR, Walker, DL. Washing mineral oil reduces contaminants and embryotoxicity. Fertil Steril. 2010; 94, 27472752.Google Scholar
234. Lee, S, Cho, M, Kim, E, et al. Renovation of a drop embryo cultures system by using refined mineral oil and the effect of glucose and/or hemoglobin added to a serum-free medium. J Vet Med Sci. 2004; 66, 6366.Google Scholar
235. Baltz, JM. Media composition: salts and osmolality. Methods Mol Biol. 2012; 912, 6180.Google Scholar
236. Moravek, M, Fisseha, S, Swain, JE. Dipeptide forms of glycine support mouse preimplantation embryo development in vitro and provide protection against high media osmolality. J Assist Reprod Genet. 2012; 29, 283290.Google Scholar
237. Xie, Y, Zhong, W, Wang, Y, et al. Using hyperosmolar stress to measure biologic and stress-activated protein kinase responses in preimplantation embryos. Mol Hum Reprod. 2007; 13, 473481.Google Scholar
238. Brinster, RL. Studies on the development of mouse embryos in vitro. I. The effect of osmolarity and hydrogen ion concentration. J Exp Zool. 1965; 158, 4957.Google Scholar
239. Hadi, T, Hammer, MA, Algire, C, Richards, T, Baltz, JM. Similar effects of osmolarity, glucose, and phosphate on cleavage past the 2-cell stage in mouse embryos from outbred and F1 hybrid females. Biol Reprod. 2005; 72, 179187.Google Scholar
240. Hammer, MA, Kolajova, M, Leveille, M, Claman, P, Baltz, JM. Glycine transport by single human and mouse embryos. Hum Reprod. 2000; 15, 419426.Google Scholar
241. Dumoulin, JC, van Wissen, LC, Menheere, PP, et al. Taurine acts as an osmolyte in human and mouse oocytes and embryos. Biol Reprod. 1997; 56, 739744.Google Scholar
242. Swain, JE, Cabrera, L, Xu, X, Smith, GD. Microdrop preparation factors influence culture-media osmolality, which can impair mouse embryo preimplantation development. Reprod Biomed Online. 2012; 24, 142147.Google Scholar
243. Swain, JE. Is there an optimal pH for culture media used in clinical IVF? Hum Reprod Update. 2012; 18, 333339.Google Scholar
244. Phillips, KP, Leveille, MC, Claman, P, Baltz, JM. Intracellular pH regulation in human preimplantation embryos. Hum Reprod. 2000; 15, 896904.Google Scholar
245. Dale, B, Menezo, Y, Cohen, J, DiMatteo, L, Wilding, M. Intracellular pH regulation in the human oocyte. Hum Reprod. 1998; 13, 964970.Google Scholar
246. Hentemann, M, Mousavi, K, Bertheussen, K. Differential pH in embryo culture. Fertil Steril. 2011; 95, 12911294.Google Scholar
247. Squirrell, JM, Lane, M, Bavister, BD. Altering intracellular pH disrupts development and cellular organization in preimplantation hamster embryos. Biol Reprod. 2001; 64, 18451854.Google Scholar
248. Zander-Fox, DL, Mitchell, M, Thompson, JG, Lane, M. Alterations in mouse embryo intracellular pH by DMO during culture impair implantation and fetal growth. Reprod Biomed Online. 2010; 21, 219229.Google Scholar
249. Hong, KH, Lee, H, Forman, EJ, Upham, KM, Scott, RT Jr. Examining the temperature of embryo culture in in vitro fertilization: a randomized controlled trial comparing traditional core temperature (37 degrees C) to a more physiologic, cooler temperature (36 degrees C). Fertil Steril. 2014; 102, 767773.Google Scholar
250. Almeida, PA, Bolton, VN. The effect of temperature fluctuations on the cytoskeletal organisation and chromosomal constitution of the human oocyte. Zygote. 1995; 3, 357365.Google Scholar
251. Pickering, SJ, Braude, PR, Johnson, MH, Cant, A, Currie, J. Transient cooling to room temperature can cause irreversible disruption of the meiotic spindle in the human oocyte. Fertil Steril. 1990; 54, 102108.Google Scholar
252. Sun, XF, Wang, WH, Keefe, DL. Overheating is detrimental to meiotic spindles within in vitro matured human oocytes. Zygote. 2004; 12, 6570.Google Scholar
253. Wang, WH, Meng, L, Hackett, RJ, Odenbourg, R, Keefe, DL. Limited recovery of meiotic spindles in living human oocytes after cooling-rewarming observed using polarized light microscopy. Hum Reprod. 2001; 16, 23742378.Google Scholar
254. Wang, WH, Meng, L, Hackett, RJ, Oldenbourg, R, Keefe, DL. Rigorous thermal control during intracytoplasmic sperm injection stabilizes the meiotic spindle and improves fertilization and pregnancy rates. Fertil Steril. 2002; 77, 12741277.Google Scholar
255. Cooke, S, Quinn, P, Kime, L, et al. Improvement in early human embryo development using new formulation sequential stage-specific culture media. Fertil Steril. 2002; 78, 12541260.Google Scholar
256. Swain, JE. Decisions for the IVF laboratory: comparative analysis of embryo culture incubators. Reprod Biomed Online. 2014; 28, 535547.Google Scholar
257. Fujiwara, M, Takahashi, K, Izuno, M, et al. Effect of micro-environment maintenance on embryo culture after in-vitro fertilization: comparison of top-load mini incubator and conventional front-load incubator. J Assist Reprod Genet. 2007; 24, 59.Google Scholar
258. Walker, MW, Butler, JM, Higdon, HL 3rd, Boone, WR. Temperature variations within and between incubators – a prospective, observational study. J Assist Reprod Genet. 2013; 30, 15831585.Google Scholar
259. Daniel, JC Jr. Cleavage of mammalian ova inhibited by visible light. Nature. 1964; 201, 316317.Google Scholar
260. Goto, Y, Noda, Y, Mori, T, Nakano, M. Increased generation of reactive oxygen species in embryos cultured in vitro. Free Radic Biol Med. 1993; 15, 6975.Google Scholar
261. Schumacher, A, Fischer, B. Influence of visible light and room temperature on cell proliferation in preimplantation rabbit embryos. J Reprod Fertil. 1988; 84, 197204.Google Scholar
262. Oh, SJ, Gong, SP, Lee, ST, Lee, EJ, Lim, JM. Light intensity and wavelength during embryo manipulation are important factors for maintaining viability of preimplantation embryos in vitro. Fertil Steril. 2007; 88, 11501157.Google Scholar
263. Takenaka, M, Horiuchi, T, Yanagimachi, R. Effects of light on development of mammalian zygotes. Proc Natl Acad Sci U S A. 2007; 104, 1428914293.Google Scholar
264. Li, R, Liu, Y, Pedersen, HS, Callesen, H. Effect of ambient light exposure of media and embryos on development and quality of porcine parthenogenetically activated embryos. Zygote. 2015; 23, 378383.Google Scholar
265. Nakayama, T, Noda, Y, Goto, Y, Mori, T. Effects of visible light and other environmental factors on the production of oxygen radicals by hamster embryos. Theriogenology. 1994; 41, 499510.Google Scholar
266. Li, R, Pedersen, KS, Liu, Y, et al. Effect of red light on the development and quality of mammalian embryos. J Assist Reprod Genet. 2014; 31, 795801.Google Scholar
267. Korhonen, K, Sjovall, S, Viitanen, J, et al. Viability of bovine embryos following exposure to the green filtered or wider bandwidth light during in vitro embryo production. Hum Reprod. 2009; 24, 308314.Google Scholar
268. Ottosen, LD, Hindkjaer, J, Ingerslev, J. Light exposure of the ovum and preimplantation embryo during ART procedures. J Assist Reprod Genet. 2007; 24, 99103.Google Scholar
269. Nakahara, T, Iwase, A, Goto, M, et al. Evaluation of the safety of time-lapse observations for human embryos. J Assist Reprod Genet. 2010; 27, 9396.Google Scholar
270. Xie, Y, Wang, F, Puscheck, EE, Rappolee, DA. Pipetting causes shear stress and elevation of phosphorylated stress-activated protein kinase/jun kinase in preimplantation embryos. Mol Reprod Dev. 2007; 74, 12871294.Google Scholar
271. Palermo, GD, Neri, QV, Rosenwaks, Z. Safety of intracytoplasmic sperm injection. Methods Mol Biol. 2014; 1154, 549562.Google Scholar
272. Mastenbroek, S, Twisk, M, van Echten-Arends, J, et al. In vitro fertilization with preimplantation genetic screening. N Engl J Med. 2007; 357, 917.Google Scholar
273. Gardner, DK, Sheehan, CB, Rienzi, L, Katz-Jaffe, M, Larman, MG. Analysis of oocyte physiology to improve cryopreservation procedures. Theriogenology. 2007; 67, 6472.Google Scholar
274. Pal, R, Mamidi, MK, Das, AK, Bhonde, R. Diverse effects of dimethyl sulfoxide (DMSO) on the differentiation potential of human embryonic stem cells. Arch Toxicol. 2012; 86, 651661.Google Scholar
275. Galvao, J, Davis, B, Tilley, M, et al. Unexpected low-dose toxicity of the universal solvent DMSO. FASEB J. 2014; 28, 13171330.Google Scholar
276. Czysz, K, Minger, S, Thomas, N. DMSO efficiently down regulates pluripotency genes in human embryonic stem cells during definitive endoderm derivation and increases the proficiency of hepatic differentiation. PLoS One. 2015; 10, e0117689.Google Scholar
277. Thaler, R, Spitzer, S, Karlic, H, Klaushofer, K, Varga, F. DMSO is a strong inducer of DNA hydroxymethylation in pre-osteoblastic MC3T3-E1 cells. Epigenetics. 2012; 7, 635651.Google Scholar
278. Shih, W, Rushford, DD, Bourne, H, et al. Factors affecting low birthweight after assisted reproduction technology: difference between transfer of fresh and cryopreserved embryos suggests an adverse effect of oocyte collection. Hum Reprod. 2008; 23, 16441653.Google Scholar
279. Luke, B, Brown, MB, Wantman, E, et al. Increased risk of large-for-gestational age birthweight in singleton siblings conceived with in vitro fertilization in frozen versus fresh cycles. J Assist Reprod Genet. 2017; 34, 191200.Google Scholar
280. Barnes, FL. The effects of the early uterine environment on the subsequent development of embryo and fetus. Theriogenology. 2000; 53, 649658.Google Scholar
281. Walker, KJ, Green, MP, Gardner, DK. Spatial asynchronous transfer of cleavage-stage mouse embryos to the uterus compromises fetal development. Mol Reprod Dev. 2015; 82, 80.Google Scholar
282. Papanikolaou, EG, Kolibianakis, EM, Tournaye, H, et al. Live birth rates after transfer of equal number of blastocysts or cleavage-stage embryos in IVF. A systematic review and meta-analysis. Hum Reprod. 2008; 23, 9199.Google Scholar
283. Glujovsky, D, Farquhar, C, Quinteiro Retamar, AM, Alvarez Sedo, CR, Blake, D. Cleavage stage versus blastocyst stage embryo transfer in assisted reproductive technology. Cochrane Database Syst Rev. 2016, CD002118.Google Scholar
284. White, CR, Denomme, MM, Tekpetey, FR, et al. High frequency of imprinted methylation errors in human preimplantation embryos. Sci Rep. 2015; 5, 17311.Google Scholar
285. Adashi, EY, Barri, PN, Berkowitz, R, et al. Infertility therapy-associated multiple pregnancies (births): an ongoing epidemic. Reprod Biomed Online. 2003; 7, 515542.Google Scholar
286. Bergh, C. Single embryo transfer: a mini-review. Hum Reprod. 2005; 20, 323327.Google Scholar